You are on page 1of 15

4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific

entific Reports

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best
experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the
meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

Article | OPEN | Published: 17 July 2017

NMR and DFT investigations of


structure of colchicine in
various solvents including
density functional theory
calculations
Gregory K. Pierens, T. K. Venkatachalam & David C. Reutens

Scientific Reportsvolume 7, Article number: 5605 (2017) | Download Citation

Abstract
A detailed NMR investigation of the chemical shifts of hydrogen and carbon atoms associated with the structure of the
naturally occurring alkaloid colchicine was conducted using high field NMR. Initially, the experimental chemical shifts for
colchicine in chloroform and DMSO were compared to the values calculated using density functional theory (DFT). There were
significant deviations observed for the chloroform solvent, but these were only slight in the DMSO solution. Dilution of the
chloroform solution changed the experimental chemical shifts and improved agreement with the DFT calculations, suggesting
self-aggregation at higher concentrations. A dimeric model was proposed for which agreement with the DFT calculated
chemical shifts was better than for corresponding monomeric structures. Three further solvents were studied to evaluate
changes in chemical shift values at different dilutions. Chloroform, benzene and water showed significant chemical shift
changes implying self-aggregation, whereas DMSO and acetone did not show significant change upon dilution.

Introduction
https://www.nature.com/articles/s41598-017-06005-5 1/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

Colchicine is a naturally occurring alkaloid that binds to tubulin, inhibiting the formation of microtubules in mitotic spindles1 and
suppressing cell division2. A number of clinical trials have been conducted utilizing colchicine as an anti-cancer drug, however
clinical use is hampered by toxicity. Although nuclear magnetic resonance (NMR) investigations of colchicine have been
conducted for nearly 30 years, there are several remaining discrepancies in the interpretation of the spectral data for this
compound. In the first report on 13C NMR chemical shifts for colchicine by Singh et al.3, assignment of the individual carbon
resonances was performed using structurally similar compounds such as tropolones. However, the low field strength of the
spectrometer used in this study (60 MHz), necessitated the use of a large amount of compound and several resonances were
superimposed, rendering assignment of the carbons difficult. Gaffied et al.4 used NMR techniques to ascertain the
conformational isomerism of the structurally related colchicine analog, isocolchicine. The absolute configuration of natural
colchicine was obtained in X-ray crystal studies by Brossi et al.5 (−) Colchicine was found to show a negative Cotton effect
curve while (+) colchicine showed the opposite; this was ascribed to a skewed phenyl substituted tropolone structure6.
Additional NMR studies were conducted by Mweksuriyen et al.7 to complete the 13C NMR assignment of the carbons in light of
discrepancies in the earlier reports8, 9 despite a large literature regarding the structural aspects of colchicine10,11,12,13,14,15. The
discrepancies associated with earlier studies can be ascribed due to the overlapping signals in the NMR spectrum as well as
the lack of advanced pulse programs for detailed NMR analysis that are now available. For example the chemical shifts have
been reported as being 150 ppm, 153.8 ppm and 151.1 ppm for C1 and 151.4 ppm and 153.5 ppm for C38, 9. Recently, Virgili et
al.16 conducted a detailed study on colchicine using multinuclear NMR. DFT calculations were performed to confirm the
chemical shifts obtained experimentally but the calculated values for certain carbon atoms differed considerably from
experimental values. The gas phase Gauge Independent Atomic Orbitals (GIAO) calculated carbon chemical shifts for several
carbons (C1, C4, C6, C18 and C9) also differed from experimental values. These discrepancies prompted us to re-examine
the proton and carbon chemical shifts for colchicine using high field NMR (700 MHz). This manuscript reports the assignments
of proton and carbon chemical shifts of colchicine in several solvents. To resolve the discrepancies in the literature, we aimed
to determine whether the polarity of solvents affected observed chemical shifts for individual carbons in the structure of
colchicine. Additionally, we examined whether the concentration of colchicine in particular solvents affected chemical shift
values.

Experimental
All chemicals were purchased from Sigma-Aldrich and used without further purification. “Concentrated” samples were
prepared as ∼46 mM of colchicine for all solvents used in this study. Additional samples were either diluted 100 fold from the
∼46 mM solution or prepared fresh as a ~0.46 mM solution. All NMR samples had volumes of 0.6 ml. For experiments using
CDCl3, the solvent was exposed to basic alumina before preparing the samples to eliminate acidity from the solvent.

NMR data were acquired on a 700 MHz Bruker Avance III HD spectrometer equipped with a TCI cryoprobe. Here chemical
shifts are reported in parts per million relative to the respective residual solvent peak as an internal standard. For the 46 mM
solutions, the 1H NMR sweep width was 9 ppm and the number of scans was 32; the 13C NMR sweep width was 200 ppm and
the number of scans was 256. Full structural elucidation was undertaken using 1H and 13C 1D NMR and a range of 2D NMR
experiments (COSY, HSQC and HMBC). The number of scans was set at 2, 4 and 8 respectively for the COSY HSQC and
HMBC datasets. The number of increments in the indirect dimension was set at 256 for all 2D experiments. All experiments
were set up using the prosol table in TOPSPIN 3.2. For the dilute samples, the scans were adjusted to give adequate signal to
noise. The DOSY experiments were acquired with the ledbpgp2s pulse sequence with δ set to 4 ms and Δ set to 100 ms.
Twenty four different gradient steps were used ranging from 2–95% with linear spacing. A recycle delay of 4 seconds was
used, which included the acquisition time. The number of scans was set to 8 and 32 for the 46 mM and 0.46 mM samples.

Molecular modeling and DFT calculations

https://www.nature.com/articles/s41598-017-06005-5 2/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

Monte Carlo Conformational searching was performed using Macromodel (Schrodinger, LLC, New York, New York, USA)17.
Torsional sampling using a Monte Carlo Multiple Minimum (MCMM) search was performed with 1000 steps per rotatable bond.
Each step was minimized with the OPLS-2005 force field using the Truncated Newton Conjugate Gradient (TNCG) method
with maximum iterations of 50,000 and energy convergence threshold of 0.02. All other parameters were left as the default
values. The lowest energy conformations (<5 kcal/mol, 50 conformations) were optimized in Gaussian18.

All conformers were optimized with Gaussian 0918 using B3LYP/631 G(d) in vacuum and the vibrational frequencies where
checked for a true minimum, i.e. no negative frequencies. All true minima were compared to remove identical structures or
conformations which where <1% of the Boltzmann population. This resulted in 8 unique conformers for colchicine. The 8
unique conformations were further optimized with B3LYP/6311 + G(2d,p) and a Polarizable Continuum Model (PCM) for
chloroform or DMSO and the vibrational frequencies where checked again for a true minimum. The free energies from the
B3LYP/6311 + G(2d,p) calculation to calculate the Boltzmann Population and used to calculate the average chemical shifts.

NMR parameters (nmr = giao) were calculated with a single-point calculation, using two functional and basis set combinations;
mpw1pw91/6-311 +G(2d,p) and B3LYP/6-311+G(2d,p)) using the optimized structures from the B3LYP/6-311 +G(2d,p)
calculation. These two functional and basis set combinations were selected due to previously calculations giving good
agreement between experimentally measured and calculated chemical shifts19,20,21. The integrated equation formalism
polarized continuum model (IEFPCM)22 for chloroform or DMSO were used in all NMR calculations and the IEFPCM has been
summarized and discussed elsewhere23,24,25. The computed NMR shielding tensors were converted to chemical shifts by the
use of empirical scaling factors23, 26 that are derived from linear regression analysis of a test set of molecules27 at the same
level of theory. The slope and intercept values of the scaling factors for mpw1pw91/6-311 + G(2d,p) and B3LYP/6-311 + 
G(2d,p) calculations are reported in the supporting information. Calculations for the dimer model used different functionals and
basis sets, described later in this manuscript.

Results and Discussion


Solutions of colchicine in deuterated chloroform and DMSO were made up to ~46 mM concentration. The proton and carbon
chemical shifts were confirmed by first principles using the 1H, 13C, COSY, HSQC and HMBC spectra. The chemical shifts
were also calculated using density functional theory (DFT) and comparisons with experimentally measured chemical shifts are
shown in Tables 1 and 2. The numbering of each of the atom in the colchicine molecule used in the present investigation is
shown in Fig. 1.

Table 1 1H Experimental and DFT calculated chemical shifts for colchicine in chloroform and DMSO
solvents using two functional and basis set combinations.

Table 2 13C experimental and DFT calculated chemical shifts for colchicine in chloroform and DMSO
solvents using two functional and basis set combinations.

https://www.nature.com/articles/s41598-017-06005-5 3/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

Figure 1

The numbering system used for colchicine (1) molecule for NMR analysis.

Two different functional and basis set combinations were used to calculate the DFT chemical shifts and it was found that the
two methods (mpw1pw91/6-311 + g(2d,p) and B3LYP/6-311 + G(2d,p)) had very similar results for chloroform and DMSO
solvents. From this point onwards the mpw1pw91/6-311 + G(2d,p) method will be discussed since it had the lowest maximum
deviation between the experimentally measured and the DFT calculated chemical shifts.

The differences between measured proton NMR chemical shifts and DFT calculated values were larger in chloroform solvent
with the largest difference being 2.55 ppm for the amide proton (Table 1). The next largest deviation was 0.38 ppm for H-8.
Other large differences were observed for protons of the C ring and the methoxy group attached to position 10. The mean
average deviation (MAE) for the chloroform comparison, calculated excluding the amide proton, was 0.19 ppm. The amide
proton was not used in the calculation of the MAE since it is possibly in exchange with water or could be involved in hydrogen
bonding interactions with itself or the solvent molecules and the chemical shift depends on strength of the hydrogen bond.

In DMSO, the maximum deviation was for the amide proton at 0.29 ppm, with H-11 and H-4 the next largest deviations at 0.30
and 0.28 ppm, respectively. The MAE for measurement made in DMSO was 0.13 ppm, also excluding the amide proton for the
reasons given above. In both solvents a good agreement (within <0.2 ppm) between experimental and calculated chemical
shifts was obtained but the DFT calculated MAE using DMSO had a slightly better fit for the whole molecule than that
calculated for the chloroform solution. Interestingly, the upfield or downfield chemical shift changes seen in the experimental
data associated with changing solvents were not always reflected by the DFT calculated chemical shifts. For example H-4 had
https://www.nature.com/articles/s41598-017-06005-5 4/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

a chemical shift of 6.52 ppm in chloroform and 6.76 ppm in DMSO and the DFT chemical shift showed a similar trend (6.41 
ppm in chloroform and 6.48 ppm in DMSO). The DFT showed a smaller change in chemical shifts with solvent than the
experimentally measured values. H-8 showed a chemical shift change from 7.57 with chloroform to 7.13 ppm with DMSO while
the DFT showed a change in the opposite direction (7.19 with chloroform to 7.25 ppm with DMSO).

The carbon chemical shift data for chloroform and DMSO showed the largest changes for C-12a, C-16(C=O), C-9(C=O) and
C-7a of >1.5 ppm. The largest changes in DFT calculated chemical shift values were for C-1, C-2, C-8, C-10 and C-12. It
should be noted that the DFT calculations were for an isolated molecule using an implicit solvent model whereas the
experimental measurements are made for compound surrounded by solvent molecules potentially with intermolecular
interactions. This limitation highlights the caution required in using these types of computer models to calculate chemical shifts
and in interpreting the results. Interestingly, the MAEs for both solvents were below 2 ppm (1.94 and 1.71 ppm for chloroform
and DMSO respectively) which represents the data is in a good agreement.

Overall the calculated chemical shifts were in slightly better agreement with experimental measurements for DMSO. However,
the large deviations for H-8 and C-9 in chloroform necessitated a detailed investigation. Two hypotheses were proposed to
explain the observed deviations. The first hypothesis was that the chloroform solvent interacts with the carbonyl at C9 and the
methoxy oxygen at C10. The second hypothesis is that colchicine molecules self-associate in solution at high concentrations.
Self-association of colchicine has been observed by Chabin et al.28 in water using NMR spectroscopy.

To test both hypotheses, the chloroform solution was diluted approximately 100 times to a concentration of ~0.46 mM. If
chloroform coordination was responsible for the observed differences, we would expect no change in the measured chemical
shifts. On the other hand, colchicine self-association was predicted to change chemical shifts. DMSO was also diluted in the
same way to test for concentration-related effects. Only 1H chemical shifts were investigated and are shown in Table 3.

Table 3 Comparison of 1H chemical shifts of colchicine in chloroform and DMSO at concentrations of 46 
mM and 0.46 mM.

As can be seen in Table 3, the chloroform solutions showed significant concentration effects but the DMSO solutions showed
extremely minimal deviations. The largest chemical shift changes were observed for H6, H8 and NH protons in chloroform.
The observed change for the NH proton chemical shift was very significant, from 7.85 ppm to 5.89 ppm upon dilution, i.e. a
change of ~2 ppm. From the observed up field shift, we infer that the amide hydrogen (NH) is involved in a much stronger
hydrogen bond at the higher concentration and upon dilution is involved in much weaker hydrogen bond at the lower
concentrations. In contrast, for studies in DMSO solution, the NH proton chemical shift did not change and remained at ∼8.56 
ppm. Therefore from these results we can conclude that colchicine in chloroform solution appear to self-aggregate at higher
concentrations due to the concentration dependence of the proton chemical shifts, i.e. H6, H8 and NH. While in DMSO
solution the molecule does not show any concentration dependence but the amide proton (NH) shows some evidence of
interaction with the solvent.

Comparing the experimental and DFT calculated chemical shifts for the concentrated and diluted solutions of colchicine in
chloroform, we observed that the diluted solution showed a reduction in the MAE from 0.19 to 0.15 ppm; the amide proton was
not used in the calculation of this metric. From Table 4 we can see that the deviations between the experimental and
calculated shifts was lower with dilution; for example the value for proton H8 decreased from 0.38 to 0.15 ppm. Based on
these observations, we can eliminate the first hypothesis, in which chloroform solvent was coordinating with colchicine.
Interestingly, the amide proton deviation decreased from 2.52 ppm to 0.56 ppm.

Table 4 Comparison of DFT calculated and experimentally measured chemical shifts at 46 mM and 0.46 
mM.

Diffusion Ordered Spectroscopy (DOSY) experiments were performed to examine differences in diffusion coefficients which
would support the self-association hypothesis. Identical experiments were run on the 46 mM and the 0.46 mM samples in
chloroform and DMSO solvents. A section of the 2D DOSY spectrum in chloroform is shown in Fig. 2. For chloroform, the
higher concentration resulted in a diffusion coefficient of 1.54 × 10−9 m2/s compared to −1.77 × 10−9 m2/s (a change of 0.32 × 
10−9 m2/s) for the diluted sample. In DMSO solvent (see supporting information), the diffusion coefficient changed by only 0.75 
× 10−10 m2/s with dilution.
https://www.nature.com/articles/s41598-017-06005-5 5/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

Figure 2

Overlay of the DOSY spectra for the 46 mM and 0.36 mM colchicine solutions in chloroform.

As a consequence of the changes in chemical shifts and diffusion coefficient values with dilution in chloroform, a DFT
calculation for a dimeric model of colchicine was performed. Using the information above, a cut-down version of colchicine (2)
was used to reduce the overall calculation time since the largest changes observed were in the C ring. Hence, the three
methoxy groups attached to the A ring of the colchicine molecule were removed. In using the cut down monomer 2 for
calculation, we took the precaution of confirming that the chemical shifts of interest were not changed significantly from those
of the parent colchicine molecule in the area of largest deviation in the experimentally measure chemical shifts between the
concentrated and diluted solutions. The cut-down monomer 2 and dimer 2 molecules are shown in Fig. 3. We emphasize that
this is a simplified dimer model as an aid to understanding the observed change in chemical shifts.

Figure 3

The partial structure of colchicine molecule used for DFT calculations.

https://www.nature.com/articles/s41598-017-06005-5 6/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

The chemical shift for the protons associated with ring C and other attached groups were compared between the full and the
cut-down colchicine monomers (Table 5). The observed changes in 1H chemical shifts were not significant, supporting the use
of the cut-down symmetric dimer for the purposes of this calculation. A symmetric dimer structure was selected since a single
set of chemical shifts was observed in the NMR data. Two hydrogen bonds were utilized to hold the dimeric structure together,
between the C9 carbonyl oxygen and the amide proton as shown in Fig. 3. This would allow the amide proton to be involved in
a hydrogen bond, in keeping with its chemical shift change upon dilution, and preferentially places H8 between two aromatic
rings.

Table 5 Comparison of key DFT calculated NMR chemical shifts between colchicine 1 and Monomer 2.

Conformational searching of monomer 2 and Dimer 2 was undertaken using Macromodel17. For monomer 2 there were 5
conformers found and the DFT calculations using Gaussian18 were conducted as describe previously. After structure
optimization removal of duplicate structures resulted in 3 conformations which were used in the DFT calculation of chemical
shifts. For the dimer 2, a weak intramolecular hydrogen bond constraint (NH-CO distance of ~2 Å) and only one conformation
was identified. Optimization with Gaussian18 was performed using wB97XD/6-311 + g(2d,p) since the functional includes
empirical dispersion forces29. No constraints were imposed for the hydrogen bond. The DFT optimized structure resulted in an
intramolecular NH–CO hydrogen bond that was 1.91 Å in length. The NMR chemical shifts were calculated for chloroform
solution using three different functional and basis set combinations to calculate the NMR chemical shifts, B3LYP/6-311 + 
g(2d,p), mpw1pw91/6-311 + g(2d,p) and wB97XD/6-311 + g(2d,p); the last two are shown in Tables 6 and 7 (also see
supporting information).

Table 6 Selected experimental and DFT calculated proton chemical shifts for Monomer 2 and Dimer 2
compared to the 46 mM colchicine in chloroform.

Table 7 Selected experimental and DFT calculated proton chemical shifts for Monomer 2 and Dimer 2
compared to the 0.46 mM colchicine in chloroform.

The DFT calculated proton chemical shifts are shown in Table 6 for Monomer 2 and Dimer 2 for selected proton as compared
to the experimental chemical shift of colchicine at 46 mM in chloroform. The results obtained from the two functional and basis
set combinations for dimer 2 were very similar. The major differences were found for H-8, H-11 and the NH which varied by
0.17, 0.12 and 0.10 ppm respectively. The experimental chemical shift for H-8 in the 46 mM sample was 7.57 ppm which lies
between the calculated values for the monomer and dimer models. This suggests that 46 mM colchicine in chloroform was
neither entirely in monomer nor in dimer form in solution. This explanation was further supported by the observed NH chemical
shift between the calculated value for the monomer and dimer. The other protons had lower calculated chemical shift values
than those measured experimentally. Therefore from these results we conclude that colchicine in chloroform solution at
concentration between 0.46 and 46 mM is a mixture of monomer and dimer.

In 1990, Chabin et al.28 reported that in water, protons in the A and C rings of colchicine moved up field with increasing
concentration and inferred that both conjugated rings were involved in self-association or dimerization. This is different to our
observations in chloroform solution where only the protons of C ring deviated significantly. Chabin et al.28 stated that the
conformation of colchicine did not change with increasing concentration.

We expanded our investigation of concentration-dependent effects by studying the proton NMR chemical shifts of several
other deuterated solvents; acetone, benzene and D2O. Figure 4 shows the differences in chemical shifts between
concentrated (~46 mM) and diluted samples (~0.46 mM) for all solvents used. Three distinct patterns emerged. Acetone and
DMSO exhibited small differences between the two concentrations. Chloroform and benzene showed an up field shift upon
dilution and water resulted in down field shifts with dilution, consistent with the values reported by Chabin et al.28.

https://www.nature.com/articles/s41598-017-06005-5 7/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

Figure 4

Differences observed in the proton chemical shift values for all the protons in various solvents of colchicine
molecule.

The results obtained for acetone and DMSO solutions suggest that these solvents interfere with the formation of the dimer. In
these solvents, the NH shift also did not change, implying that acetone and DMSO solvents themselves interacts with the
amide proton, possibly forming a hydrogen bond since the shift remained at approximately 7.7 ppm. For chloroform and
benzene, the major changes were observed mainly in the C ring, H6 and the acetamide group. Behavior in benzene and in
chloroform solution appeared similar, pointing to self-aggregation in both solvents. In benzene solution, the amide proton also
showed a large change with dilution (~3.4 ppm); in dilute solutions the amide proton is unlikely to be involved in a hydrogen
bond. Changes with dilution were larger in benzene than in chloroform but both were in an up field direction.

For solutions in water, dilution resulted in a downfield shift and most of the protons affected are in A, B and C rings, as
reported by Chabin et al.28. The largest changes observed were for H5, H11 and H12 protons. This may be explained by self-
aggregation that is different from that in chloroform and benzene solutions since the H8 proton is not among the most affected
chemical shifts.

In conclusion we have observed that colchicine appears to self-associate in solution in chloroform, benzene and water but not
in DMSO or acetone. Self-association in benzene and chloroform seems to principally involve the C ring with the amide proton
moving up field with dilution. In water it seems to involve A, B and C rings. The amide proton of colchicine in DMSO and
acetone did not change in their chemical shift values with dilution, remaining >7.5 ppm, implying that they may be hydrogen
bonded to the solvent.

https://www.nature.com/articles/s41598-017-06005-5 8/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

Additional information
Publisher's note: Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional
affiliations.

References
Soifer, D. & Mack, K. In Structural Elements of the Nervous System 245–280 (Springer US, 1984).

Brossi, A. Bioactive alkaloids. 4. Results of recent investigations with colchicine and physostigmine. J. Med. Chem 33, 2311–
2319 (1990).

Singh, S., Parmar, S., Stenberg, V. & Farnum, S. Carbon-13 nuclear magnetic resonance spectrum of colchicine. Spectrosc.
Lett. 10, 1001–1012 (1977).

Gaffield, W., Lundin, R. E. & Horowitz, R. M. Conformational isomerism and its relation to the mutarotation of isocolchicine.
Chem. Commun. 610–612 (1984).

Brossi, A. et al. aS, 7S‐absolute configuration of natural (−)‐colchicine and allocongeners. FEBS Lett 262, 5–7 (1990).

Hrbek, J. et al. Circular dichroism of alkaloids of colchicine type and their derivatives. Collect. Czech. Chem. Commun. 47,
2258–2279 (1982).

Meksuriyen, D., Lin, L.-J., Cordell, G. A., Mukhopadhyay, S. & Banerjee, S. K. Nmr Studies of Colchicine and its
Photoisomers, β-and λ-Lumicolchicines. J. Nat. Prod. 51, 88–93 (1988).

Danieli, B., Palmisano, G. & Ricca, G. S. C-13 NMR analysis of colchicine and isocolchicine-a revision of colchicine
assignments. Gazz. Chim. Ital. 110, 351–352 (1980).

Elguero, J., Muller, R., Blade‐Font, A., Faure, R. & Vincent, E. Carbon‐13 Magnetic Resonance Spectroscopy. A Study Of
Colchicine And Related Compounds. Bull. Soc. Chim. Belg. 89, 193–204 (1980).

Allen, F. H. The Cambridge Structural Database: a quarter of a million crystal structures and rising. Acta Crystallogr. Sect. B:
Struct. Sci. 58, 380–388 (2002).

Allen, F. H. & Motherwell, W. S. Applications of the Cambridge Structural Database in organic chemistry and crystal chemistry.
Acta Crystallogr. Sect. B: Struct. Sci 58, 407–422 (2002).

Blade-Font, A., Muller, R., Elguero, J., Faure, R. & Vincent, E.-J. Carbon-13 Nuclear Magnetic Resonance Spectrum of
Colchicine: A Reassignment. Chem. Lett. 8, 233–236 (1979).

He, H.-P., Liu, F.-C., Hu, U. & Zhu, H.-Y. Alkaloids from the Flowers of Colchicum automnale. Acta. Bot. Yunnanica 21, 364
(1999).

Hufford, C. D., Capraro, H. G. & Brossi, A. 13C‐and 1H‐NMR. Assignments for colchicine derivatives. Helv. Chim. Acta. 63,
50–56 (1980).

Lessinger, L. & Margulis, T. The crystal structure of colchicine. A new application of magic integers to multiple-solution direct
methods. Acta Crystallogr. B 34, 578–584 (1978).

https://www.nature.com/articles/s41598-017-06005-5 9/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

Virgili, A. et al. A Spectroscopic study of colchicine in the solid state and in solution by multinuclear magnetic resonance and
vibrational circular dichroism. Helv. Chim. Acta. 97, 471–490 (2014).

MacroModel, version 10.4, Schrödinger, LLC, New York, NY, 2014.

Gaussian 09, Revision D.01, M. J. Frisch et al., Gaussian, Inc., Wallingford CT, 2013.

Fraser, J. A., Lambert, L. K., Pierens, G. K., Bernhardt, P. V. & Garson, M. J. Secondary metabolites of the sponge-derived
fungus Acremonium persicinum. J. Nat. Prod. 76, 1432–1440 (2013).

White, A. M. et al. Rearranged diterpenes and norditerpenes from three Australian Goniobranchus mollusks. J. Nat. Prod. 79,
477–483 (2015).

Pierens, G. K., Venkatachalam, T. & Reutens, D. C. Comparison of experimental and DFT‐calculated NMR chemical shifts of
2‐amino and 2‐hydroxyl substituted phenyl benzimidazoles, benzoxazoles and benzothiazoles in four solvents using the IEF‐
PCM solvation model. Magn. Reson. Chem (2015).

Tomasi, J., Mennucci, B. & Cances, E. The IEF version of the PCM solvation method: an overview of a new method addressed
to study molecular solutes at the QM ab initio level. Journal of Molecular Structure: THEOCHEM 464, 211–226 (1999).

Lodewyk, M. W., Siebert, M. R. & Tantillo, D. J. Computational prediction of 1H and 13C chemical shifts: A useful tool for
natural product, mechanistic, and synthetic organic chemistry. Chem. Rev. 112, 1839–1862 (2011).

Wiitala, K. W., Al‐Rashid, Z. F., Dvornikovs, V., Hoye, T. R. & Cramer, C. J. Evaluation of various DFT protocols for computing
1H and 13C chemical shifts to distinguish stereoisomers: diastereomeric 2‐, 3‐, and 4‐methylcyclohexanols as a test set. J.
Phys. Org. Chem. 20, 345–354 (2007).

Rablen, P. R., Pearlman, S. A. & Finkbiner, J. A comparison of density functional methods for the estimation of proton
chemical shifts with chemical accuracy. The Journal of Physical Chemistry A 103, 7357–7363 (1999).

Willoughby, P. H., Jansma, M. J. & Hoye, T. R. A guide to small-molecule structure assignment through computation of (1H
and 13C) NMR chemical shifts. Nat. Protoc. 9, 643–660 (2014).

Pierens, G. K. 1H and 13C NMR scaling factors for the calculation of chemical shifts in commonly used solvents using density
functional theory. J. Comput. Chem. 35, 1388–1394 (2014).

Chabin, R. M., Feliciano, F. & Hastie, S. B. Effect of tubulin binding and self-association on the near-ultraviolet circular dichroic
spectra of colchicine and analogs. Biochemistry 29, 1869–1875 (1990).

Chai, J.-D. & Head-Gordon, M. Long-range corrected hybrid density functionals with damped atom–atom dispersion
corrections. PCCP 10, 6615–6620 (2008).

Acknowledgements
The research was funded in part by funding from the Australian Research Council.

Author information

https://www.nature.com/articles/s41598-017-06005-5 10/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

Author notes
Gregory K. Pierens and T. K. Venkatachalam contributed equally for this work.

Affiliations

The Centre for Advanced Imaging, The University of


Queensland, Building 57, Research Road, St. Lucia,
Queensland, 4072, Australia

Gregory K. Pierens, T. K. Venkatachalam & David C. Reutens

Contributions
G.K.P.: Co-developed the initial idea, performed NMR and computational studies, analyzed NMR and computational results
and wrote manuscript. T.V.: Co-developed the initial idea, performed NMR studies, analyzed NMR results and wrote
manuscript. D.C.R.: Aided with manuscript writing and editing.

Competing Interests
The authors declare that they have no competing interests.

Corresponding author
Correspondence to T. K. Venkatachalam.

Electronic supplementary material


Supporting information - https://static-
content.springer.com/esm/art%3A10.1038%2Fs41598-

https://www.nature.com/articles/s41598-017-06005-5 11/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

017-06005-
5/MediaObjects/41598_2017_6005_MOESM1_ESM.pdf

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use,
sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the
original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The
images or other third party material in this article are included in the article’s Creative Commons license, unless indicated
otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended
use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the
copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.

About this article


Publication history

Received

14 December 2016

Accepted

05 June 2017

Published

17 July 2017

DOI
https://doi.org/10.1038/s41598-017-06005-5

https://www.nature.com/articles/s41598-017-06005-5 12/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

Share this article


Anyone you share the following link with will be able to read this content:

Get shareable link

Subjects
Density functional theoryStructure elucidation

Further reading
Friedel-Crafts Reaction of N,N-Dimethylaniline with
Alkenes Catalyzed by Cyclic Diaminocarbene-Gold(I)
Complex - https://doi.org/10.1038/s41598-018-29854-0
Hangzhi Wu, Tianxiang Zhao & Xingbang Hu

Scientific Reports (2018)

Comments
By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that
does not comply with our terms or guidelines please flag it as inappropriate.

https://www.nature.com/articles/s41598-017-06005-5 13/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

0 Comments Scientific Reports  Login

 Recommend t Tweet f Share Sort by Newest

Start the discussion…

LOG IN WITH
OR SIGN UP WITH DISQUS ?

Name

Email

Password

I agree to Disqus' Terms of Service

I agree to Disqus' processing of email and IP address, and the use of cookies, to
facilitate my authentication and posting of comments, explained further in the
Privacy Policy

Be the first to comment.

✉ Subscribe d Add Disqus to your siteAdd DisqusAdd 🔒 Disqus' Privacy PolicyPrivacy PolicyPrivacy

Scientific Reports
ISSN 2045-2322 (online)

About us
Press releases
Press office
Contact us

https://www.nature.com/articles/s41598-017-06005-5 14/15
4/6/2019 NMR and DFT investigations of structure of colchicine in various solvents including density functional theory calculations | Scientific Reports

https://www.nature.com/articles/s41598-017-06005-5 15/15

You might also like